An improved solvent-free synthesis of flunixin and 2-(arylamino) nicotinic acid derivatives using boric acid as catalyst

pdf
Số trang An improved solvent-free synthesis of flunixin and 2-(arylamino) nicotinic acid derivatives using boric acid as catalyst 10 Cỡ tệp An improved solvent-free synthesis of flunixin and 2-(arylamino) nicotinic acid derivatives using boric acid as catalyst 1 MB Lượt tải An improved solvent-free synthesis of flunixin and 2-(arylamino) nicotinic acid derivatives using boric acid as catalyst 0 Lượt đọc An improved solvent-free synthesis of flunixin and 2-(arylamino) nicotinic acid derivatives using boric acid as catalyst 1
Đánh giá An improved solvent-free synthesis of flunixin and 2-(arylamino) nicotinic acid derivatives using boric acid as catalyst
4.9 ( 11 lượt)
Nhấn vào bên dưới để tải tài liệu
Để tải xuống xem đầy đủ hãy nhấn vào bên trên
Chủ đề liên quan

Nội dung

Yarhosseini et al. Chemistry Central Journal (2017) 11:124 https://doi.org/10.1186/s13065-017-0355-4 Open Access RESEARCH ARTICLE An improved solvent‑free synthesis of flunixin and 2‑(arylamino) nicotinic acid derivatives using boric acid as catalyst Mahsa Yarhosseini, Shahrzad Javanshir* , Zahra Dolatkhah and Mohammad G. Dekamin Abstract A simple solvent-free protocol for the preparation of flunixin, a potent non-narcotic, non-steroidal anti-inflammatory drugs is reported using boric acid as catalyst. Its salt, flunixin meglumine are then prepared under reflux in EtOH. This sustainable method are then extended for the synthesis of a series of 2-(arylamino) nicotinic acid derivatives. The present protocol combines non-hazardous neat conditions with associated benefits like excellent yield, straightforward workup, and use of readily available and safe catalyst in the absence of any solvent, which are important factors in the pharmaceutical industry. The pathway for catalytic activation of 2-chloronicotic acid with boric acid was also investigated using Gaussian 03 program package. Keywords: Homogeneous catalysis, Solvent-free, Non-steroidal anti-inflammatory drugs (NSAID), N-Methyl-dglucamine, Boric acid, Anilino-nicotinic acid Introduction Flunixin, 2-(2-methyl-3-trifluoromethylanilino)nicotinic acid (Fig. 1) is a non-narcotic, non-steroidal anti-inflammatory drug (NSAID) and cyclooxygenase inhibitor with potent antipyretic and analgesic activity, parenterally administered in clinical and veterinary applications [1– 3]. It is frequently administered in the form of the meglumine salt. The analgesic activity of N-methyl-d-glucamine salt of flunixin (flunixin meglumine) is comparable to that observed by powerful narcotic analgesics such as morphine, meperidine hydrochloride, and pentazocine. However, the undesirable side effects are less compared to the above mentioned narcotics analgesics and drug dependence liability does not occur [4]. Apart from its parenteral analgesic activities, flunixin meglumine also possesses anti-inflammatory attributes [5–8]. Flunixin meglumine was first reported in 1975 [9] and is the active pharmaceutical ingredient in numerous drug products such as *Correspondence: shjavan@iust.ac.ir Heterocyclic Chemistry Research Laboratory, Department of Chemistry, Iran University of Science and Technology, Tehran 16846‑13114, Iran Resflor, Banamine Solution, Banamine Paste, Banamine Granules and Finadyne [10–19]. So far, several improved methods have been developed for the synthesis of flunixin including classical reflux in water [12], xylene [5] and ethylene glycol [20]. Some of these methods suffer from several drawbacks such as difficult workup, long reaction times, use of large quantities of non-green organic solvents such as xylene that are harmful for environment and harsh reaction conditions. In this regard, solvent-free approach which has become increasingly popular in recent years for reasons of economy and pollution prevention, as well as cleaner products, simple work-up, high-speed due to the high concentration of materials, and excellent yields would be the ideal approach, as it is often claimed that “the best solvent is no solvent” [21]. To the best of our knowledge, we herein report the first green synthesis of 2-(2-methyl-3-trifluoromethylanilino)nicotinic acid (3a) under solvent-free conditions in the presence of catalytic amount of Boric acid. In this contribution, studies to reach a broader scope and generality of this reaction is also presented using other anilines, as nucleophiles, some 2-arylaminonicotinic acids derivatives 3 were also synthesized by this method. © The Author(s) 2017. This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/ publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated. Yarhosseini et al. Chemistry Central Journal (2017) 11:124 Fig. 1 Chemical structure of flunixin Results and discussion Boric acid [­H3BO3 or B(OH)3] has attracted particular attention in recent years as catalyst in organic synthesis [22–25], because of many advantages such as uncomplicated handling, inexpensiveness, eco-friendly nature and commercially available [26]. Consequently, we have considered the synthesis of 2-(2-methyl-3-trifluoromethylanilino)nicotinic acid (3a), in the presence of boric acid under solvent-free conditions, and its salt, flunixin meglumine (6) under reflux in EtOH (Scheme 1). In a first attempt, the reaction of 2-methyl-3-trifluoromethylanilin (1) with 2-chloronicotinic acid (2) was examined under various reaction conditions, the results and the Scheme 1 Synthesis of flunixin meglumine Page 2 of 10 optimization of reaction conditions are summarized in Table 1. Initially, the effect of acid and basic catalysts was tested. In the absence of any catalyst no reaction took place under reflux in water (Table 1, entry 1), and also under solvent-free condition the desired product was obtained in low yield in the absence of any catalyst (entry 2). When this reaction is carried out in the presence of a catalyst such as ­K2CO3, ­NEt3, ­Fe3O4, and DABCO in the presence and absence of solvent only a trace amount of product was detected even after 24 h (entries 5–8), while the desired product was obtained in moderate yield in the presence of PTSA and boric acid (entries 9 and 10). Considering the cost and toxicity of these two catalysts, and also due to a slight difference in efficiency between them (entries 21 and 23), we choose boric acid as optimal catalyst. In the next step the effect of various solvents and temperature (80–150 °C) was examined (Table 1, entries 8–22). The best yield was obtained at 120 °C under solvent-free condition (entry 17). Therefore, considering the viewpoints of green chemistry, the synthesis of flunixin was carried out under solvent-free condition. And finally, we studied the effect of the ratio of reactants on the yield of the reaction. When the reagents were used in an Yarhosseini et al. Chemistry Central Journal (2017) 11:124 Page 3 of 10 Table 1 Optimization of the reaction conditions for the synthesis of (3a) Entry Ratio of 1/2 Catalyst Catalyst loading (mole %) Solvent Temp (°C) Time (h) Yieldc (%) 1 2:1 – – H2O Reflux 24 – 2 2:1 – – – 100 10 35 3 2:1 K2CO3 15 – 100 24 Trace 4 2:1 NEt3 15 – 100 24 – 5 2:1 K2CO3 15 H2O Reflux 24 Trace 6 2:1 NEt3 15 H2O Reflux 24 – 7 2:1 Fe3O4 15 H2O Reflux 24 – 8 2:1 DABCOa 15 H2O Reflux 24 Trace 9 2:1 PTSA 15 H2O Reflux 24 65 10 2:1 H3BO3 15 H2O Reflux 24 60 11 2:1 H3BO3 15 EtOH Reflux 24 40 12 2:1 H3BO3 15 n-Hexanol Reflux 24 62 13 2:1 H3BO3 15 PEG-400b Reflux 24 30 14 2:1 H3BO3 15 DMF Reflux 24 45 15 2:1 H3BO3 15 Xylene Reflux 24 Trace 14 2:1 H3BO3 15 Toluene Reflux 24 50 17 2:1 H3BO3 15 – 100 10 68 18 2:1 H3BO3 30 – 100 10 82 19 2:1 H3BO3 48 – 100 10 84 20 2:1 H3BO3 30 – 80 12 45 21 2:1 H3BO3 30 – 120 10 90 22 2:1 H3BO3 30 – 150 10 90 23 2:1 PTSA 30 – 120 9 90 24 1:1 H3BO3 30 – 120 20 60 Reaction conditions: 2-methyl-3-trifluoromethylanilin (2 mmol), 2-chloronicotinic acid (1 mmol) The optimum reaction conditions are in italics a 1,4-diazabicyclo[2.2.2]octane b Polyethylene glycol c The yields refer to the isolated product equimolar ratios, a reduction in yield was observed (entry 24) and the optimum ratio of 2-methyl-3-trifluoromethylanilin (1) to 2-chloronicotinic acid (2) was 2:1 (entry 21). Therefore, the optimum reaction conditions for the synthesis of flunixin were as follows: molar ratio of 2-methyl-3-trifluoromethylanilin (1) to 2-chloronicotinic acid (2) equal to 2:1 at 120 °C under solvent-free condition in the presence of ­H3BO3 (20 mg, 30 mol%) as catalyst (Table 1, entry 21). This procedure was also scaled up to 30 g of product and the yield of flunixin was CF3 COOH CH3 NH2 44.45g (0.254 mol) N Cl 20g (0.127 mol) Scheme 2 Scalability of the reaction to the multi-gram scale maintained excellent however the reaction time was augmented about 45 min (Scheme 2). The synthesized flunixin was characterized by FT-IR, 1 HNMR, 13CNMR and GC/MS (see Additional file 1). A quantitative analysis was performed to determine the amount of residual boron in the synthesized flunixin. The result of ICP analysis shows a level of 2.23 ppm (2.23 mg/L) of boron in the product. According to the World Health Organization, the health-based guideline for boron level in drinking water is 2.4 mg/L. Boric acid (30 mol %) 3a Solvent free, 120oC, 10h 45min 32 g, 85% Yarhosseini et al. Chemistry Central Journal (2017) 11:124 Page 4 of 10 Table 2 Results of the synthesis of 2-arylaminonicotinic acids derivatives 3(b–l) under solvent-free conditions COOH N R N H 3(b-l) Entry Anilines 1 2-Chloroaniline 2 2,6-Dichloroaniline 3 3,4-Dichloroaniline 4 2,3-Dichloroaniline 5 2,4-Dimethylaniline 6 3-Nitroaniline 7 4-Nitroaniline 8 1-Naphthylamine 9 2-Aminophenol Products Time (min) Yielda (%) Mp (°C) 20 62 213–215 [212–214] [28] 60 25 264–267 [266–267] [29] 25 72 258–260 [260] [30] 25 58 256 [256–258] [5] 10 92 225–227 [225] [31] 30 90 218–219 [216–218] [32] 60 28 274–275 [272–274] [29] 30 85 191–194 [193–194] [33] 25 88 230–232 [230–232] [34] 3b 3c 3d 3e 3f 3g 3h 3i 3j Yarhosseini et al. Chemistry Central Journal (2017) 11:124 Page 5 of 10 Table 2 continued COOH N R N H 3(b-l) Entry Anilines 10 Methylamine 11 Buthylamine 12 Phenylethylamine Products Time (min) Yielda (%) Mp (°C) – – – – – – – – [228–230] [35] 3k 3l 3m Reaction conditions: amines (2 mmol), nicotinic acid (1 mmol), and 20 mg ­H3BO3 at 120 °C under solvent-free condition a The yields refer to the isolated product Since we have been very successful in this solvent free approach with a very specific aromatic amine such as 2-methyl-3-trifluoromethylanilin, we decided to study the scope and limitation of the reaction using other aniline derivatives as nucleophiles. Reactions of aniline derivatives with 2-chloronicotinic acid gave good-to-excellent yields. This result was expected because the pyridine ring is activated toward nucleophilic attack by the formation of a pyridinium salt with the acid catalyst. Table 2 shows several 2-arylaminonicotinic acids derivatives synthesized under solvent-free condition in the presence of 20 mg ­H3BO3 at 120 °C. As it can be seen, nucleophilic substitution takes place readily at the 2- and 4-positions of a ring particularly when substituted with an effective leaving group such as a halogen atom. In nucleophilic substitution, the intermediate is negatively charged. The capacity of the ring to withstand the negative charge determines the stability of the intermediate and transition state which drives it and consequently determines the reaction rate. Nucleophilic attack at the 2-position yields a carbanion that is hybrid of structures. The hybrid structures of the anionic intermediate formed during the nucleophilic attack are especially stable, since the negative charge is located on the electronegative nitrogen atom which can better accommodate it (Scheme 3). For this reason the nucleophilic substitution occurs specially on the pyridine ring instead of the benzene ring and preferably at the 2 and 4 positions. Various anilines were tolerated in this methodology and gave the product in good yield (Table 2, entries 1, 3, 4, 5, 6, 8, 9), but some anilines (entry 2 and 7) evince the influence of electron withdrawing group and steric hindrance on their reactivity by giving a low product yield. Remarkably, when primary amines other than anilines were selected for this transformation, no reaction was observed and they are not potent nucleophiles in this methodology; this could be attribute to the higher nucleophilicity of the acyclic primary amines poisoning OCOOH N Cl COOH Nu OH N Cl Nu N Cl Meisenheimer complex Scheme 3 Nucleophilic substitution of a pyridine ring COOH N Cl Nu Yarhosseini et al. Chemistry Central Journal (2017) 11:124 Page 6 of 10 the catalyst as a result of competitive binding with the primary amine to the boric acid. The scope in substrates for nucleophilic substitution was also investigated using some pyridines derivatives such as 4-chloro pyridine, 3-chloro pyridine and 2-chloro-5-nitro pyridine. However, when the reaction was run using these substrates, no product was observed even after 12 h (Table 3). The Table 3 Scope in pyridine analogues for nucleophilic aromatic substitution Entry Substrate Nucleophile Time (h) Yield (%) 1 12 – 2 12 – 3 12 – Reaction conditions: aniline (2 mmol), pyridine analogues (1 mmol), and 20 mg (0.3 mmol) ­H3BO3 at 120 °C under solvent-free condition effect of other substituents on the pyridine ring show that the substitution reactions are more facile if an activating group is present on the ring. 2-chloronicotinic acid contains a strongly electron-withdrawing group –COOH. This would be expected to better activate the halogen located ortho to it. This result clearly show the activating effect of the carboxylic acid group in S ­ NAr of aromatic ring by forming resonance-stabilized intermediate known as Meisenheimer complex [27]. The results for synthesis of 2-arylaminonicotinic acids derivatives 3 are very satisfactory, affording the respective products in good yield in short reaction times. Flunixin is not soluble in water, and as it is administered by intravenous or intramuscular injection consequently to increase its solubility in water, it is often formulated as the meglumine salt. Therefore the reaction of flunixin (1 mmol) with meglumine (1 mmol) was also examined under various reaction conditions and the results and the optimization of reaction conditions are summarized in Fig. 2. As it can be seen, refluxing in ethanol gave the best results among all examined conditions. The as produced flunixin meglumine was characterized by FTIR, 1HNMR, Opmizaon condions for the synthesis of (6) 130 Yield* 98 100 Time 2 0 2 0 Ethanol 60 60 60 Iso - propanol 25 Acetonitrile 20 60 78.4 82.6 60 80 82 Temp - Ethanol - Fig. 2 Optimization conditions for the synthesis of (6). Reaction conditions: flunixin (1 mmol), meglumine (1 mmol). *The yields refer to the isolated product a b OH HO B N OH Cl COOH Fig. 3 Two possible pathway for catalytic activation of 2-chloronicotic acid with Boric acid. a HBD, b Boron-Nitrogen Lewis acid–base interaction Yarhosseini et al. Chemistry Central Journal (2017) 11:124 Page 7 of 10 a charge distribution IMA PA b IMB1 IMB 81 PB TS Fig. 4 The optimized structures of nicotinic acid-boric acid, IMs, and product for pathway (a, b) by B3lyp/6-311G (d, p) singlet in the gas phase Yarhosseini et al. Chemistry Central Journal (2017) 11:124 Page 8 of 10 13 CNMR, GC/Mass (available in Additional file 1) and high-performance liquid chromatographic (HPLC) method was used to determine the purity to flunixin meglumine (see Additional file 1). As can be seen from the HPLC chromatogram, the purity of the synthesized flunixin meglumine is higher than the standard. Theoretical study of mechanism The pathway for catalytic activation of 2-chloronicotic acid with Boric acid was investigated using theoretical gas-phase calculations [36]. The full geometry optimizations and property calculations were performed within the density functional theory (DFT) approach using the Becke’s three-parameter B3LYP exchange–correlation functional [37] and the 6-311G** basis set [38, 39]. The two possible pathway for catalytic activation of 2-chloronicotic acid with Boric acid, i.e. hydrogen bonding and lone pair co-ordinate bond with the empty orbital on the boron are presented in Fig. 3. The optimized geometries of IMs, products (P) and transition state (TS) and the numbering used in the analysis of the results are displayed in Fig. 4. The optimized bond lengths (in nm) and stability energies E (a.u.) for IMs, TS and products are tabulated in Table 4. The calculations overall indicate small affinity of boron for the nitrogen donor atom (IMB1) so that the optimized structure tend to form hydrogen bonds (HBDs) in (IMB 81) and give strong support to the suggested mechanism in which HBD formation between nitrogen in pyridine and OH of Boric acid is the preferred mode of activation. These results also showed that the internuclear distance C1–Cl10 increased from 1.75 in IMA to 2.04 in TS indicating the cleavage of this bonds while the inter-nuclear distance N22–C1 decreased from 1.91 in Table 4 The optimized bond lengths (in nm) and stability energies E (a.u.) computed at B3LYP/6-311G (d, p) level of theory for IMA, IMB, TS, PA, and PB (see Fig. 4 for atoms numbering) Bond length (nm) IMA IMB (N6–H19) 1.92 1.95 (C1–Cl10) 1.75 1.75 − 1149.199 a.u. − 1149.193 a.u. (C1–N22) E (stability energies in a.u.) TS PB 1.92 1.92 1.37 1.37 − 1352.544 a.u. − 1352.544 a.u. 2.04 1.91 Scheme 4 Plausible mechanisms for the synthesis of flunixin PA Yarhosseini et al. Chemistry Central Journal (2017) 11:124 TS to 1.37 (N21–C1) in PA. These observations provide evidence that the reaction proceeds through an addition elimination mechanism. We also used Gaussian 03 [36] to observe the vibrations in the TS and saw that the negative frequency corresponded to the motion of N22 and Cl10. According to obtained results, the following proposed mechanism seems reasonable for the synthesis of flunixin catalyzed by boric acid (Scheme 4). The substitution reactions proceed via an addition–elimination mechanism. In this proposed mechanism the nucleophilic aromatic substitution reaction occurs via a two-step mechanism in which the first step is the addition 2-methyl-3-trifluoromethylanilin (1) to the pyridine ring of 2-chloronicotinic acid (2) which was activated through HBD between pyridine and Boric acid and so leading to the intermediate (I). In the second step elimination of the halide ion restores the aromaticity leading to the product (3a). Conclusion In summary, we have developed a simple, convenient and efficient method for the synthesis of flunixin using ­H3BO3 as catalyst under solvent-free conditions. This method are then extended for the synthesis of a series of 2-(arylamino)nicotinic acid derivatives. The present protocol has several advantages, particularly solvent- free conditions, high yields, ecofriendly operational, experimental simplicity along with mild conditions and the most important of them is development and optimization of flunixin meglumine synthesis in order to transfer to a larger scale for manufacture. Density function UB3LYP/6-311++g(d,p) calculations give strong support to the suggested mechanism in which HBD formation between nitrogen in pyridine and OH of Boric acid is the preferred mode of activation. Additional file Additional file 1. Supporting Information. Authors’ contributions SJ have designed the study, participated in discussing the result, and revised the manuscript. MY carried the literature study, performed the assays, conducted the optimization as well as purification of compounds, and prepared the manuscript. ZD performed the NMR analyzes and assay validation studies. MGD is the second advisor and participate in final sequence alignment. All authors read and approved the final manuscript. Acknowledgements The authors gratefully acknowledge the support of the Research Council of the Iran University of Science and Technology, Tehran, IRAN. We also would like to express our very great appreciation to Mr. Khorsandi for his valuable time and help. Page 9 of 10 Associated content Experimental procedure and product characterization data: IR, 1H NMR, 13C NMR, GC/Mass and HPLC of compounds 3a and 6 are reported in Additional file. Competing interests The authors confirm that there are no known conflicts of interest associated with this publication and there has been no significant financial support for this work that could have influenced its outcome. Ethics approval and consent to participate Not applicable. Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. Received: 9 October 2017 Accepted: 21 November 2017 References 1. Doran HJ (1993) United States Patent 07, 812, 183 2. Houdeshell JW, Hennessey PW (1977) A new nonsteroidal, anti-inflammatory analgesic for horses. J Equine Med Surg 1:57 3. Tobin T (1979) Pharmacology review: the nonsteroidal anti-inflammatory drugs. II. Equiproxen, meclofenamic acid, flunixin and others [used in horses]. J Equine Med Surg 3:298 4. Ciofalo VB, Latranyi MB, Patel JB, Taber RI (1977) Flunixin meglumine: a non-narcotic analgesic. J Pharmacol Exp Ther 200:501 5. McKillop A (1993) United States Patent 5, 248, 781 6. Jang D, Park J, Ko H, Jee J, Kim J, Kim S, Kim C, Jang DJ, Park JS, Ko HR, Jee JP, Kim JK, Kim ST, Kim CK (2005) Simultaneous determination of niflumic acid and its prodrug, talniflumate in human plasma by high performance liquid chromatography. Biomed Chromatogr 19:32 7. Cocco MT, Congiu C, Onnis V, Morelli M, Felipo V, Auli O (2004) Synthesis of new 2-arylamino-6-trifluoromethylpyridine-3-carboxylic acid derivatives and investigation of their analgesic activity. Bioorg Med Chem 12:4169 8. Yasunobu N, Toshikazu K (1990) Japanese Patent 02, 45, 451 9. Sherlock MH (1975) United States Patent 3, 891, 761 10. Towson JC, Coveney D (2013) United States Patent 8, 501, 788 B2 11. Gutsche CD, Johnson HE (1955) Ring enlargements. III. Ring enlargement of cyclohexanone with ethyl N-nitroso-N-benzylcarbamates carrying methyl and methoxyl substituents on the phenyl nucleus. J Am Chem Soc 77:109–112 12. Doran HJ, Coveney DJ (1995) European Patent 0, 619, 809 13. Sherlock MH, Long JF (1980) United States Patent 4, 205, 073 14. Hongbin Y, Na W (2006) Chinese Patent CN1803773 A 15. Sherlock MH (1974) United States Patent Office 3, 839, 344 16. Martin S, Wong YS (1980) United States Patent 4, 172, 095 17. Martin S, Wong YS (1980) United States Patent 4, 209, 464 18. Takeshi K, Toshikazu K, Hideki O (1988) Japanese Patent 86, 188, 576 19. Lamendola JF, Dhiru V, Robert T (1988) European Patent 295, 604 20. Jaouhari R, Quinn P, Jaouhari R, Quinn P (1994) Improved process for the preparation of 2 methyl-3-trifluoromethylaniline: a versatile intermediate for flunixin synthesis. Heterocycles 38:2243 21. Tanaka K (2009) Solvent-free organic synthesis, 2nd edn. Wiley-VCH, Weinheim 22. Nguyen TB, Jonathan S, Minh QT, Ludmila E, Ali M (2012) Boric acid: a highly efficient catalyst for transamidation of carboxamides with amines. Org Lett 14:3202 23. Ravi KM, Kondaiah G, Naveenkumar K, Raju V, Purandhar K, Apurba B, Rakeshwar B (2007) Boric acid catalyzed amidation in the synthesis of active pharmaceutical ingredients. Org Process Res Dev 11:1065 24. Thomas S, Hansen A, Jerrik M, Anders R (2011) Synergy of boric acid and added salts in the catalytic dehydration of hexoses to 5-hydroxymethylfurfural in water. Green Chem 13:109 Yarhosseini et al. Chemistry Central Journal (2017) 11:124 25. Tang PW (2005) Boric acid catalyzed amide formation from carboxylic acids and amines: N-benzyl-4-phenylbutyramide. Org Synth 81:262 26. Shelke KF, Sapkal SB, Kakade GK, Shinde PV, Shingate BB, Shingare MS (2009) Boric acid as an efficient catalyst for the synthesis of 1, 1-diacetate under solvent-free condition. Chin Chem Lett 20:1453 27. Meisenheimer J (1902) Ueber reactionen aromatischer nitrokörper. Eur J Org Chem 323(2):205 28. Li ZH, Xia ZN, Chen G (2011) Catalyst-free amination of 2-chloronicotinic acid in water under microwave irradiation. J Chem Res 35:709 29. Stampa A (1973) France Patent 2236862-A1 30. Mefetah H, Brouant P, Galy A, Galy J, Barbe J (1995) Potential anticancer benzo-naphthyridones with fused rings: a theoretical model for predicting orientation in the cyclization of intermediates. Med Chem Res 5:522 31. Montginoul B, Giral T (1980) Synthese de nouvelles 1H, 3H pyrido[2,3-d] pyrimidinediones-2,4. J Heterocycl Chem 17:235 32. Kolodziejczyk A, Arendt A (1975) Synthesis of a tumor-growth inhibitor-dihydrochloride of 1-nitro-9-/dimethylamino-propylamino/acridine/C-283/tritium labelled in the acridine ring and 14C-labelled in the side chain. J Label Compd 11:385 33. Li Z, Xiao Sh, Liang R, Xia Z (2012) Synthesis of 2-(arylamino) nicotinic acids in high-temperature water. Res Chem Intermed 38:1691 34. Sherlock M, Kaminski J, Tom C, Lee F (1988) Antiallergy agents. 1. Substituted 1, 8-naphthyridin-2 (1H)-ones as inhibitors of SRS-A release. J Med Chem 31:2108 Page 10 of 10 35. Scherrer RA, Beatty HR (1980) Preparation of o-substituted benzoic acids by the copper(II)-catalyzed reaction of diphenyliodonium-2-carboxylate with anilines and other nucleophiles. J Org Chem 45:2127 36. Gaussian 03, Revision C 02, Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Zakrzewski VG, Montgomery JA, Stratmann RE, Burant JC, Dapprich S, Millam JM, Daniels AD, Kudin KN, Strain MC, Farkas O, Tomasi J, Barone V, Cossi M, Cammi R, Mennucci B, Pomelli C, Adamo C, Clifford S, Ochterski J, Petersson GA, Ayala PY, Cui Q, Morokuma K, Malick DK, Rabuck AD, Raghavachari K, Foresman JB, Cioslowski J, Ortiz JV, Baboul AG, Stefanov BB, Liu G, Liashenko A, Piskorz P, Komaromi I, Gomperts R, Martin RL, Fox DJ, Keith T, Al-Laham MA, Peng CY, Nanayakkara A, Challacombe M, Gill PMW, Johnson B, Chen W, Wong MW, Andres J, Gonzalez C, Head-Gordon M, Replogle ES, Pople JA (1998) Gaussian 98, Revision A.9. Gaussian Inc, Pittsburgh 37. Becke AD (1988) Density-functional exchange-energy approximation with correct asymptotic behavior. Phys Rev A 38:3098 38. Lee CT, Yang WT, Parr RG (1988) Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys Rev B 37:785 39. Frisch M, Head-Gordon M, Pople J (1990) Semi-direct algorithms for the MP2 energy and gradient. Chem Phys Lett 166:281
This site is protected by reCAPTCHA and the Google Privacy Policy and Terms of Service apply.